Zheng Chen and Yacine Chitour

11Controllability of Keplerian motionwith low-thrust control systems

Zheng Chen, Math. Lab. of Orsay, Univ. Paris-Sud, CNRS, University Paris-Saclay, 91400 Orsay, France, [email protected]

Yacine Chitour, L2S-Supélec, University Paris-Saclay & CNRS, 91192 Gif-sur-Yvette, France,[email protected]

Abstract: In this paper, we present the controllability properties of Keplerian motion controlled by low-thrust control systems. A low-thrust control system, compared with a high or an even impulsive control system, provides a fuel-efficient means to control the Keplerian motion of a satellite in the restricted two-body problem. We obtain that, for any positive value of maximum thrust, the motion is controllable for orbital transfer problems. For two other typical problemsde-orbit problem and orbital insertion problem, which have state constraints the motion is controllable if and only if the maximum thrust is bigger than a limiting value. Finally, two numerical examples are given to show the numerical method to compute the limiting value.

Keywords: Keplerian motion, low thrust, controllability

AMS classification: 49K15, 70Q05

11.1Introduction

In classical mechanics, the determination of the motion of two celestial bodies, which interact only with each other, is the typical two-body problem. If one body is light enough, the uncontrolled motion of the light body around a heavy body is a restricted two-body problem, and the motion is well known as Keplerian motion. A common example is the artificial bodies (i.e., spacecrafts or satellites) moving around the Earth. Once the atmospheric effects are negligible and the Earth overwhelmingly dominates the gravitational influence, a satellite moves stably on a periodic orbit if the mechanical energy of the satellite is negative. As an increasing number of artificial satellites have been launched into space around the Earth or even into deeper space since the mid of the last century, an important problem arises in astronautics, that is, to control the Keplerian motion of a satellite to transfer between different orbits to achieve desiredmission requirements. The control of an artificial satellite is generally performed by system propulsion, expelling mass in a high speed to generate an opposite reaction force according to Newtons third law of motion. Up to now, there are already several types of propulsion systems available, including chemical propulsion systems and electric propulsion systems. Though chemical propulsion systems are able to provide much higher thrust, electric propulsion systems have the potential for a much higher specific impulse than is available from chemical ones, resulting in a lower fuel consumption and thus a longer satellite lifetime for a given propellant mass. On the one hand, the electric propulsion systems provide a fuel-efficient means to control the motion of a satellite. On the other hand, the possible maximum thrust provided by the electric propulsion systems is very low. As a consequence, the transfer time is exponentially long and the optimization of the transfer time is an important challenge which has been studied in [4, 5, 7]. Though the low-thrust control systems provide a fuel-efficient means to control the motion of a satellite, the problem to know whether or not it has the ability to move a satellite from one point to another one arises. This is actually a controllability property, which is a prerequisite to analyze mission feasibility during the design of a space mission or of an optimal trajectory. Restricting the mechanical energy of a satellite into a negative region without any other state constraints, the controlled motion is called the orbital transfer problem (OTP), and the controllability for OTP was derived in [5, 7] to show that there exist admissible controlled trajectories for every OTP if the maximum thrust is positive. In the current paper, the controllability for OTP is established using alternative techniques from geometric control (cf. [9, 12]). Taking into account the state constraint that the radius of a satellite is larger than the radius of the surface of the atmosphere around the Earth, the orbital insertion problem (OIP) and de-orbit problem (DOP) are defined in this paper. Some controllability properties for the OIP and DOP are then addressed and we show that there exist admissible controlled trajectories for OIPs and DOPs if and only if the maximum thrust is bigger than a specific value (depending on the initial point or final point).

The organization of the paper is as follows. In Section 11.2, we recall the basic properties of the dynamics of motion of a satellite around the Earth, and basic notations and definitions are given which are crucial for the analysis of controllability properties. In Section 11.3, the controllability for OTPs is first addressed using geometric control technology in [9]. Then, the controllability properties for OIPs and DOPs are derived in Subsections 11.3.2 and 11.3.3, respectively. In Section 11.4, two numerical examples are given to show the development in this paper. Finally, a conclusion is given in Section 11.5.

11.2Notations and definitions

11.2.1Dynamics

Consider a satellite as a mass point moving around the Earth, its state in a geocentric inertial cartesian coordinate (GICC), illustrated in Figure 1, consists of its position vector r 3{0}, velocity vector υ 3, and mass Then, the dynamics for themotion of the satellite for positive times can be written as

Fig. 1: Geocentric inertial cartesian coordinate.

where μ > 0 is the gravitational constant, β > 0 is a scalar constant determined by the specific impulse of the low-thrust control system equipped on the satellite, denotes the Euclidean norm, and the thrust (or control) vector τ 3 takes values in the admissible set

where τmax is a positive constant.

If ? = 3{0} × 3 and x = (r, υ), we define two vector fields f 0, f 1 on ? by

with 6×3 denotes the set of 6 ×3 matriceswih real entries and I 3 is the identity matrix of 3. Let ε be the closed ball in 3 centered at the origin and of radius ε > 0. For every ε > 0, we consider the control-affine system Σε given by

where the control vector u 3 takes values in ε.Wewill use in this paper the vector field point of view of [11, 13, 14]. For every point x ? and every u ε, we denote by

where f 0 and f 1 are referred to as the drift vector field and the control vector field, respectively. Note that trajectories of Σε starting at any x0 ? and associated with a measurable u : + ε are well defined on an open interval of + containing 0, which depends in general on x0 and u().

11.2.2Study of the drift vector field in X

In this section, we recall the main properties of the drift vector field f 0. For every x ?, we use γx to denote the restriction to + of the maximal trajectory of f 0 starting at x, that is, γx is defined on some interval [0, tf (x)), where tf (x) . Then, the following holds true.

Property 2.1 (First integrals [2, 6]). For every x ?, if γx(t) = (r̃(t), ̃(t)) on [0, tf (x)), the quantities

are constant along γx and the corresponding constant values are the angular momentum vector h 3, the Laplace vector L 3, and the mechanical energy of a unit mass E , which is the sum of the relative kinetic energy ̃(t) 2 /2 and the potential energy μ/ r̃(t) .

Combining equations (2.7) and (2.8), we have the following two properties:

Property 2.2 (Straight line [2, 6]). Let x ? with h = 0, that is, r and υ are collinear. Then, the trajectory γx is a straight line and tf (x) is either finite or infinite depending on initial conditions.

Property 2.3 (Conic section [2, 6]). Let x ? with h ≠ 0, that is, r and υ are not collinear. Then, tf (x) is infinite and the trajectory γx is a periodic trajectory with locus defining a conic section lying in a two-dimensional plane perpendicular to h called the orbital plane.

Let

Then define on ? the function e : x L /μ. Along every trajectory of f 0 starting at x ?̃ , one gets, after multiplying equation (2.8) by r̃(t), that

where the angle θ(t, x) is defined by Note that the previous formula holds true if L = 0, since in that case, e(x) cos θ(t, x) is equal to zero and the orbit is a circle.

Notice from equation (2.11) that an orbit (r̃(t), ̃(t)) = γx(t), with x ?̃ on + being a parabola if e(x) = 1 and a hyperbola if e(x) > 1. Put a satellite on a parabolic or ahyperbolic orbit without any control; then, it can escape to infinity: Thus, parabolic and hyperbolic orbits are generally used for a satellite to escape from the gravitational attraction of the Earth. For every point x ?̃ , if 0 e(x) < 1, the orbit γx(t) on + is an ellipse, whose orientation is illustrated in Figure 2. Moreover, it is easy to deduce the following characterization of elliptic orbits.

Fig. 2: The orientation of a two-dimensional orbital plane in GICC and the geometric shape and orientation of an elliptic orbit on the orbital plane.

Property 2.4. Given every point x ?̃ , the mechanical energy E is negative if and only if e(x) < 1.

Thus, let us define the set

then for every point (r, υ) ?, the associated orbit γx on + is periodic and the set ? is called the periodic region in ?.

Definition 2.5 (Smallest period tp). Given every point x ?, we denote by

tp : ?, x tp(x) ,

the smallest period of the orbit γx on +.

According to equation (2.11), for every point x ?, if e(x) ≠ 0, the associated orbit γx on [0, tp(x)] has its perigee point and apogee point at θ(t, x) = 0 and π, respectively. Thus, let

where rp(x) and ra(x) are the perigee and apogee distances of the orbit γx on [0, tp(x)], respectively, if e(x) ≠ 0. Note that ra(x) = rp(x) if and only if e(x) = 0, which corresponds to a circular orbit.

Property 2.6 (Minimum radius and maximum radius). Given every periodic orbit (r̃(t), ̃(t)) = γx(t) on [0, tp(x)] in ?, we have rp(x) r̃(t) ra(x) on [0, tp(x)]. Thus, the perigee distance rp(x) and apogee distance ra(x) are the minimum radius and maximum radius of the orbit (r̃(t), ̃(t)) on [0, tp(x)].

11.2.3Admissible controlled trajectory of Σsat

For every initial point and measurable control function τ() taking values in ?(τmax) with τmax > 0, we use Γ(t, τ(t), yi) to denote the restriction to + of the solution of Σsat starting from yi. Let tf̃ + be the maximum time such that the solution Γ(t, τ, yi) lies in and set ?Γ = [0, tf̃ ). Then the restriction of Γ(t, τ, yi) to ?Γ is said to be the controlled trajectory of Σsat starting from yi and associated with τ().

Remark 2.7. For every point , let (x(t), m(t)) = Γ(t, 0, yi) on ?Γ ; then, we have x(t) = γxi (t) and m(t) = mi for every t 0, that is, tf̃ = .

Definition 2.8 (Controlled Keplerian motion). Given every initial point yi = (xi , mi) ?×andmeasurable control function τ() taking values in ?(τmax) with τmax > 0, the corresponding trajectory Γ(t, τ(), yi) of Σsat is called a controlled Keplerian motion.

Let rc > 0 and M0 > 0 denote the radius of the surface of atmosphere around the Earth and the mass of a satellite without any fuel, respectively; then, given every point on the trajectories of Keplerian motions, it is required that r > rc and m > M 0.

Definition 2.9 (Admissible region). We define the set

the admissible region in ? for Keplerian motion and/ or controlled Keplerian motion.

Definition 2.10 (Admissible controlled trajectory). Given every M0 > 0, we say thatthe controlled trajectory (x(t), m(t)) = Γ(t, τ, xi , mi) of Σsat on some finite intervals[0, tf] ?Γ with the initial condition is an admissible controlledtrajectory if (x(t)) ? and m(t) M0 for t [0, tf ].

For every time interval [0, tf ] ?Γ , since ṁ (t) 0, it follows m(tf ) m(t) on [0, tf ]. Thus, the inequality m(t) M0 can be ensured by m(tf) M0.

11.2.4Controlled problems in A

For x ?, let (r̃(t), ̃(t)) = γx(t) on +; then, we have that the inequality r̃(t) > rc is satisfied on + if rp(x) > rc. Thus, we define the set

It is immediate to see that the periodic uncontrolled trajectory γ(t, x) starting at any x ?+ remains in ?+.

Let

Then, for every point x ?, there exists an interval [t1, t2] [0, tp(x)] such that r̃(t) rc for t [t1, t2]. Thus, placing a satellite on a point x ?, it can move out of the admissible region ?.

Definition 2.11 (Stable periodic region ?+ and unstable periodic region ? in ?). Wesay that the two sets ?+ and ? are the stable and unstable periodic regions, respectively.

All the satellites periodically moving around the Earth are located in the stable periodic region ?+. In order to fulfill observation or other mission requirements, a satellite is controlled to move from one point xi in ?+ to another point xf in ?+ by its control system.

Definition 2.12 (Orbital transfer problem (OTP)). We say that the problem of controlling a satellite from a point xi in ?+ to another point xf in ?+ is the OTP; see Figure 3a.

For a typical space mission, in order to place a satellite into a stable orbit in ?+, a rocket is used to carry the satellite from the surface of the Earth to a point xi in ?, at which the rocket and the satellite are separated. From this moment on, the satellite is controlled by its own control system to be inserted into a stable orbit in ?+.

Definition 2.13 (Orbital insertion problem (OIP)). We say that the problem of controlling a satellite from an initial point xi ? to a final point xf ?+ is the orbit insertion problem; see Figure 3b.

After a satellite in the stable region ?+ finishes its mission, it should be decelerated to return to the unstable region ?. Then, the satellite will coast into atmosphere such that the aerodynamic pressure will act as a control to control the satellite to fly to landing sites.

Definition 2.14 (De-orbit problem (DOP)). We say that the problem of controlling a satellite from an initial point xi ?+ to a final point xf ? is the de-orbit problem; see Figure 3c.

Fig. 3: OTP, OIP, and DOP.

11.3Controllability

According to the definition for controlled Keplerian motion (Definition 2.8), the controllability of Keplerian motion deals with the existence of admissible controlled trajectories for OTP, OIP, and DOP.

Definition 3.1 (Controllability for OTP). We say that the system Σsat is controllable for OTP if there exists τmax > 0 so that, for every initial mass mi > 0, and every initial and final points (xi , xf) (?+)2, there exists a time tf ?Γ and an admissible controlled trajectory (x(t), m(t)) = Γ(t, τ, xi , mi) of Σsat on [0, tf ] in such that x(tf) = xf .

Definition 3.2 (Controllability for OIP and DOP). We say that the system Σsat is controllable for OIP (DOP respectively) from any point xi ? (xi ?+ respectively) if for every initial mass mi > 0 there exists τmax > 0 so that, for every final point xf ?+ (xf ? respectively), there exists a time tf ?Γ and an admissible controlled trajectory (x(t), m(t)) = Γ(t, τ, xi , mi) of Σsat on [0, tf ] in such that x(tf) = xf .

For every initial point xi ? and ε > 0, we use Γε(t, u(t), xi) to denote the trajectory of Σε in equation (2.5) associated with a measurable control u() : [0, tf̄ ] ε, and we define tf̄ + as the maximum time such that Γε(t, u(t), xi) lies in ? on [0, tf̄ ). Set ?̄ = [0, tf̄ ). We refer to Γε(t, u(t), xi) as the controlled trajectory of Σε starting from xi and corresponding to the control u().

Remark 3.3. Since γx(t) = Γε(t, 0, x) on ?,̄ the uncontrolled trajectory Γε(t, 0, x) is periodic on + if x ?.

Lemma 3.4. Fix ε > 0 and Then, given every measurable controlu() : [0, tf] ε, if τmax εmi, then there exists M0 > 0 and an admissible controlled trajectory (x(t), m(t)) = Γ(t, τ, yi) of Σsat on [0, tf ] in ? × [M0, mi] such that Γε(t, u, xi) = x(t) for every t [0, tf ] and m(tf) M0.

Proof. Since m(t) mi for each time t [0, tf ] and τmax εmi, it follows that the thrust vector τ() on [0, tf ] can take values in the set ? in equation (2.2) such that τ(t)/m(t) = u(t) for every time t [0, tf ]. Thus, let (x(t), m(t)) = Γ(t, τ, xi , mi). Then, we have Γε(t, u, xi) = x(t) for every time t [0, tf ]. Since along the trajectory (x(t), m(t)) = Γ(t, τ, xi , mi) on [0, tf ], we have u(t) = τ(t)/m(t), which implies that ṁ (t) = β u(t) m(t). Thus, we obtain

Let M0 := mieβεtf > 0. Then, m(tf) M0, and the lemma is proved.

In order to study controllability, it is necessary to first show that the admissible region ? is a connected subset of ?.

Lemma 3.5 (Connectedness of ?). The admissible region ? is an arc-connected subset of ?, that is, for every initial point xi ? and every final point xf ?, there exists a continuous path f : [0, 1] ?, λ x(λ) such that x(0) = xi and x(1) = xf .

Proof. To prove that ? is arc-connected, we use the MEOE coordinates (cf. Definition 6.2). Let us choose two points zi and zf in ? given by

zi = (Pi , exi , eyi , hxi , hyi , li), zf = (Pf , exf , eyf , hxf , hyf , lf ) ,

with xi = x(zi) and xf = x(zf ). We thus define the path z : [0, 1] ? by z(λ) =(P(λ), ex (λ), ey(λ), hx (λ), hy (λ), l(λ)), where

where rf = rf and ri = ri . Note that ex(λ)2 + ey(λ)2 < 1 for each λ [0, 1]. Let g(λ) = (P(λ), ex (λ), ey (λ), hx(λ), hy (λ), l(λ)) on [0, 1]; we then have g(0) = zi and g(1) = zf . Consider the continuous function x(λ) = (r(z(λ)), υ(z(λ))) for λ [0, 1]. Itfollows that x(0) = xi and x(1) = xf. It immediately follows that x(λ) ?for λ [0, 1]. Finally, since we haver(λ) × υ(λ) ≠ 0 and on [0, 1]. This proves the lemma.

We also need the following lemma.

Lemma 3.6 (Connectedness of ?+). The set ?+ is a connected subset of ?.

Proof. Given every two points xi ?+ and xf ?+, using the same technique as in the proof of Lemma 3.5, let x(λ) = (r(z(λ)), υ(z(λ))) on [0, 1]. But we rewrite P(λ) in the following form:

where rpi = rp(xi) and rpf = rp(xf ). Then, we have

Thus, x() takes values in ?+ and this proves the lemma.

11.3.1Controllability for OTP

In this subsection, we first give a controllability property of Σε for OTP; then, according to Lemma 3.4, we will establish the controllability of Σsat for OTP.

Definition 3.7. For every controlled trajectory x̄() = Γε(, ū , xi) of Σε (where ε > 0, ū() : [0, tf ] ε is measurable and xi ?), we define

(x̄) : λ̇(t) = A(t)λ(t) + B(t)u(t) ,

the linearized system along x̄() of Σε on [0, tf ], where

A(t) = f x(x̄(t), ū(t)), B(t) = f u(x̄(t), ū(t)) ,

on [0, tf ].

We first have a result of local controllability for the systems Σεs around the periodic trajectories of the drift vector field.

Lemma 3.8. Let x̄ ?+. Then, for every ρ > 0, there exists σ > 0 such that the following properties hold: σ(x̄) ?+ and, for every x σ(x̄), there exists a controlled trajectory Γε(t, u(t), x̄) of Σε such that

Γε(0, u, x̄) = x̄, Γε(tp(x̄), u, x̄) = x ,

and

Γε(t, u, x̄) Γε(t, 0, x̄) < ρ ,

for t [0, tp(x̄)].

Proof. According to Theorem 7 of Chapter 3 in [12], it suffices to prove the controllability of the linearized system along the periodic trajectory Γε(t, 0, x̄) on the interval [0, tp(x̄)]. Then, the latter controllability would follow. According to Corollary 3.5.18 of Chapter 3 in [12], by the following rank condition, there exists a time τ [0, tp(x̄)] and a nonnegative integer k such that the rank of the matrix [B0(τ), B1(τ), . . . , Bk(τ)] equals 6, where for i = 1, 2, . . ., and B0(t) = B(t). It, therefore, amounts to compute some Bi()s. The explicit expressions for matrices A and B in terms of x are

Since B0(t) = B(t), it follows that

Thus, we have that the rank of the matrix equal to 6 forevery time t [0, tp(x̄)], proving the lemma.

Proposition 3.9. For ε > 0, the control system Σε is controllable for OTP within ?+, that is, for every initial point xi ?+ and final point xf ?+, there exists a controlled trajectory Γε(t, u, xi) of Σε in ?+ on a finite interval [0, tf ] ?̄ such that Γε(tf , u, xi) =xf .

Proof. Since the subset ?+ is path-connected as is shown by Proposition 3.6, it followsthat any two different points xi and xf in ?+ are connected by a path x : [0, 1] ?+, λ x(λ) such that x(0) = xi and x(1) = xf . By compactness of the support of x() in ?+, there exists, for every σ > 0, a finite sequence of points x0, x1, . . . , xN, on the support of x(λ) so that x0 = xi, xn = xf and xj+1 σ(xj), for j = 0, 1, . . . , N 1. According to Lemma 3.8, for every ρ > 0, there exists σ > 0 small enough and a finite sequence of points x0, x1, . . . , xN as above such that, for j = 0, 1, . . . , N 1, xj+1 σ(xj) and one has a controlled trajectory Γε(t, uj , xj) on the interval [0, tp(xj)] such that

Γε(0, uj , xj) = xj, Γε(tp(xj), uj, xj) = xj+1 ,

and

Γε(t, uj(t), xj) Γε(t, 0, xj) < ρ, for t [0, tp(xj)] .

For σ > 0, let ?j ?+ be an open neighborhood of xj such that σ(xj) ?j forj = 0,1, . . . , N and set = {Γε(t, 0,?j ), t 0}. For ρ > 0 small enough, the openset is included in ?+. By concatenating the Γε(, uj , xj), for j = 0,1, . . . , N 1,the initial point xi can be steered to xf , proving the proposition.

According to Lemma 3.4, and recalling the definition of controllability for OTPs in Definition 2.12, we obtain the following result of controllability.

Corollary 3.10. For every μ > 0, β > 0, τmax > 0, the system Σsat is controllable for OTP.

Since ?+ ?, the system Σsat is controllable for OTPs within ? for every positive τmax if the satellite takes high enough percent of total fuel, that is (mi mf )/mi > 0 is big enough. This result makes senses in engineering for electric thrust systems whose maximum thrust τmax is very small.

11.3.2Controllability for OIP

We provide next a controllability criterion for OIP.

Lemma 3.11. Assume that, for every point there exists τ > 0 and a positive time t ̄ ?Γ , and a control τ̃() : [0, t ?(τ) such that along the controlled trajectory (x̃, m̃ (t)) = Γ(t, τ̃(t), xi , mi) on [0, t, we have x̃(t) ? on [0, t, m̃ (t)̄ > 0, and rp(x̃(t)̄ ) > rc. Then, the system Σsat is controllable for OIP from (xi , mi).Proof. Note that the assumption implies that there exists a control τ̃() : [0, t ?(τ) such that the admissible controlled trajectory (x̃(t), m̃ (t)) = Γ(t, τ, xi , mi) in ? × [m( t, ̄ mi] on [0, t] ̄ steers (xi , mi) in to some (x( t)̄ , m( t)) ̄ in After arriving at x(t)̄ in ?+, according to Proposition 3.9, it follows that there exists an M0 (0, m(t)̄ ], a finite time tf ?Γ , and a control τ() : [0, tf ] ?(τ) such that along the controlled trajectory (x(t), m(t)) = Γ(t, τ, x̃(τ), m̃ (τ)) on [0, tf ], we have x(t) ? on [0, tf ], x(tf) = xf, and m(tf) > M0.

One cannot have controllability for OIP for every value of τmax > 0. Indeed, pick a point (xi , mi) in For every control τ() taking values in ?(τmax), the corresponding controlled trajectory (x(), m()) = Γ(, τ, xi , mi) converges to Γ(, 0, xi , mi) on [0, tp(xi)] as τmax tends to zero. Then, since rp(xi) < rc, there exists t [0, tp(xi)] such that r(t) < rc implying that (x(), m()) = Γ(, τ, xi , mi) is not admissible for every control τ() taking values in ?(τmax). For large values of τmax, we can steer (xi , mi) to (xf , mf) as described in the following lemma.

Lemma 3.12. For every β > 0, μ > 0, and point yi = (xi , mi) in there exists τmax such that the following holds:

(1) if τ > τmax, there exists a control τ() taking values in ?(τ) and a positive time t ̄ ?Γ such that, along the controlled trajectory (x(t), m(t)) = Γ(t, τ, yi , mi) on [0, t, we have x(t) ? on [0, t, m(t)̄ > 0, and rp(x(t)̄ ) > rc;

(2) if τ τmax, for every control τ() taking values in ?(τ), the controlled trajectory (x(), m()) = Γ(, τ, yi, mi) does not reach

Proof. In the light of the remark preceding the lemma, it is enough to find a value of τ̄ and a control τ() taking values in ?(τ̄) steering yi = (xi , mi) to some point in along an admissible trajectory. We proceed as follows. Let C0 = ri1/2υiwhich belongs to (0, Choose now the function υ() defined on some time interval [0, T] ̄ (with T ̄ to be fixed later) as follows: υ(0) =i, ṙ(t) = υ(t), and

Here r(t) denotes a continuous choice of vector perpendicular to r(t) in the 2D planespanned by ri and υi. (Implicitly, we assume with no loss of generality that υi ≠ 0 with ai > 0 and bi ≠ 0.) For simplicity, we assume next that ai = bi = 1. The curve (r(), υ()) defined previously can be explicitely integrated using polar coordinates for r() = r() exp(iθ()). One gets that, on [0, T],

After integration, one has for t [0, T],

One also checks that with eibeing a constant vector of unit normcollinear to ri × υi and One deduces that r0p2(t) = C1r(t)1/2 for some positive constant C1. One thus fixes T so that rp(T̄) > rc. It remains to determine τ̄ so that (r(t), υ(t)) is part of a controlled admissible trajectory of the system Σε. One first integrates over [0, T] the differential equation

and then takes for t [0, T]. The final bound τ̄ is simply the maximum of τ(t) over [0, T].

As a combination of Lemmas 3.11 and 3.12, we obtain the following result.

Corollary 3.13. For every β > 0, μ > 0, and initial point (xi , mi) there exists a limiting value τmax > 0 depending on (xi , mi) such that the following properties hold:

(1) if τ > τmax, the system Σsat is controllable for the OIP and

(2) if τ τmax, the system Σsat is not controllable for the OIP.

The limiting value τmax can be computed by combining a shooting method and a bisection method as described in Section 11.4.

11.3.3Controllability for DOP

Let us define a system Σ̃sat associated with Σsat as

where all variables are the same as defined in equation (2.1). For every and measurable control function τ taking values in ?(τmax) with τmax, we define by Γ̃(t, τ(t), yi) the corresponding trajectory of Σ̃sat for some positive times.

Remark 3.14. For every controlled trajectory (r(t), υ(t), m(t)) = Γ(t, τ(t), yi) of the system Σsat on some finite intervals [0, tf] ?Γ with (rf , υf , mf) = Γ(tf , τ(tf ), yi), the trajectory (r̃(t), ̃(t), m̃ (t)) = Γ̃(t, τ(tf t), rf , υf , mf ) of Σ̃sat runs backward in time along the trajectory Γ(t, τ(t), yi) on [0, tf ], that is,

As a consequence, according to Lemma 3.12 and Corollary 3.13,we obtain the following result.

Corollary 3.15. For each β > 0, μ > 0, and mi > 0, given a point (rf , υf) ?, there exists a τmax > 0 depending on (rf , υf ) and mi such that the following properties hold:

(1) if τ > τmax, the system Σsat is controllable for the corresponding DOP and

(2) if τ τmax, the system Σsat is not controllable for the corresponding DOP.

11.4Numerical examples

In this section, we consider two numerical examples, one OIP and one DOP, to compute the limiting value τmax in Corollaries 3.13 and 3.15, respectively. The gravitational constant μ in system Σsat (and/or Σ̃sat) is 3,986,000.47 km3/s2, the radius of the Earth is re = 6,374,000 m, and we consider that the vertical depth of atmosphere around the Earth is 90,000 m, which means rc = re + 90,000 m.

11.4.1A numerical example for OIP

In order to be able to compute the limiting value τmax in Corollary 3.13, we first define the following optimal control problem (OCP).

Definition 4.1 (Optimal control problem (OCP) for OIP). Given every initial point (xi , mi) and τ > 0, the OCP for OIP consists of steering a satellite by τ() ?(τ) on a time interval [0, tf] ?Γ along the system Σsat such that, along the controlled trajectory (r(t), υ(t), m(t)) = Γ(t, τ(t), xi , mi), the time tf is the first occurrence for r(tf) = rp(r(tf ), υ(tf )), that is, r(t) > rp(r(t), υ(t)) on [0, tf ), and that rp(r(tf ), υ(tf )) is maximized, that is, the cost functional is

Let tf̃ >0 be the optimal final time of the OCP for OIP and (x̃(t), m̃ (t)) = Γ(t, τ̃(t), xi , mi) on [0, tf̃ ] be the optimal controlled trajectory with the associated optimal control τ̃(t) ?(τ) on [0, tf̃ ]. One can check, by using the Pontryagin maximum principle as was done in [7], that τ̃(t) = τ on the whole interval [0, tf̃ ]. Thus, fixing the initial point (xi , mi) , we have that the final time tf̃ , the trajectory (x̃(t), m̃ (t)) at each time t [0, tf̃ ], and the final perigee distance rp(x̃(tf̃ )) are functions of τ. Thus, let us define a function

If one can find τmax > 0 such that s(τmax) = 0, then τmax is the limiting value in Corollary 3.13. For every τ > 0, using a shooting method as an inner loop to solve the OCP for OIP, we can obtain a value for s(τ). Then, using a bisection method as an outer loop, one can obtain τmax > 0 such that s(τmax) = 0. According to equation (2.14) and the objective of the OCP for OIP, place a satellite with the initial mass mi > 0 on a point (ri , υi) ?. The optimal controlled trajectory lies on a 2D plane spanned by ri and υi. Hence, the limiting value τmax in Corollary 3.13 is determined only by ri , υi , and by the flight path angle ηi [π/2, π/2], that is, the angle between the velocity vector υi and the local horizontal plane, defined by

Assume that a rocket carries a satellite whose initial mass is mi = 150 kg, from the surface of the Earth to a point xi = (ri , υi) in the unstable region ? such that ri = re + 110,000 m, υi = 7879.5 m/s, and ηi = 5. The rocket and the satellite are separated at this point xi. Then, the satellite has to use its own engine to steer itself from the point xi into the stable region ?+. We can see from Figure 4 that the periodic orbit γxi has collisions with the surface of the atmosphere around the Earth.

We choose the specific impulse of the engine fixed on the satellite as Isp= 2000 s, which implies that where g0 = 9.8 m2/s. The computed result of the limiting value is τmax = 8.052 N. Thus, in order to be able to insert the satellite from the point xi into the stable region ?+, the maximum thrust of the engine has to be larger than 8.052 N. The optimal controlled trajectory of the corresponding OCP for OIP with τ = 8.052 is plotted in Figure 4 as well. To see the numerical resultsfor different initial points, other two points x1 = (r1, υ1) and x2 = (r2, υ2) are chosen on the periodic orbit γxi such that

Fig. 4: The periodic orbital γxi and the optimal controlled trajectory of the OCP for the OIP with τ = τmax starting from xi .

r1 = re + 379,494 m, υ1 = 7,562 m/s, η1 = 4.3517° ,

r2 = re + 599,351 m, υ2 = 7,312m/s, η2 = 3.0132° .

Then, the limiting value of τmax corresponding to the two initial points xj, j = 1, 2, are computed as 9.037 N and 10.719 N, respectively.We see that the limiting values τmax are different for different initial points on the same periodic orbit. The time history of radius r(t) and the perigee distance rp(x(t)) along the optimal controlled trajectories starting from the initial points xi and xj, j = 1, 2, are plotted in Figure 5. Since the three points xi and xj, j = 1, 2 lie on the same periodic orbit γxi , rp at the initial time is the same, as shown in Figure 5.

11.4.2A numerical example for DOP

A DOP is a powered flight phase of a satellite in the region , during which a decelerating manoeuvre is performed so that the satellite will move to the desired final point xf = (rf , υf) ? at the entry interface (EI). The condition at EI permits thesatellite to have a subsequent safe entry flight in atmosphere to a landing site.Atypical condition at EI, see [1], is given as

Fig. 5: The profile of r and rp with respect to time for three different initial points, that is, xi and xj (j = 1, 2), for OIP. The subscripts, i, 1, and 2 correspond to the initial points xi , x1, and x2, respectively.

where rEI = re + 122,000 m, VEI = 7879.5m/s, and ηEI = 15° denote the norm of position vector, the norm of velocity vector, and the flight path angle at EI, respectively. In order to compute the limiting value τmax in Corollary 3.15 for the DOP to a point (rf , υf ) in ?, we first define the below OCP.

Definition 4.2 (OCP for DOP). Given every final point xf = (rf , υf) ? and τ > 0, let mi > 0 be the initial mass of a satellite, the OCP for DOP consists of steering the satellite by τ() ?(τ) on a time interval [0, tf ] ?Γ subject to the system Σ̃sat such that, along the controlled trajectory (r(t), υ(t), m(t)) = Γ̃(t, τ(t), rf , υf , mf ) (mf > 0 is free) of the System Σ̃sat, the time tf is the first occurrence for r(tf ) = rp(r(tf ), υ(tf )), mi = m(tf ), and rp(r(tf ), υ(tf )) is maximized, that is, the cost functional is the same as equation (4.1).

Given every initial mass mi > 0 and final point (rf , υf ) in ?, let tf̄ > 0 be the optimal final time of the OCP for DOP and (x̄(t), m̄ (t)) = Γ̃(t, τ̄(t), rf , υf , mf ) be the optimal controlled trajectory associated with the control τ̄(t) ?(τ) on [0, tf̄ ]. Then, the same as the OCP for OIP, the perigee distance rp(x̄(tf̄ )) is a function of τ. Let us define a function

Then, according to equation (3.3), in order to compute the limiting value τmax in Corollary 3.15, it suffices to combine a shooting method and a bisection method to compute the value τmax such that s̄(τmax) = 0.

What we developed in this paper is applicable not only for low-thrust control systems but also for high-thrust control systems if only the thrust is finite instead of impulsive. Thus, we consider the space shuttles parameters in [3, 10]. The initial mass is 95,254.38 kg. The specific impulse of the engine is 313 s, which means β = 3.26 × 104. The numerical result is τmax = 14,004.62N. Note that the propulsion for a space shuttle is provided by the orbital manoeuvring system (OMS) engines which produce a total vacuum thrust of 53,378.6 N; see [3, 10]. Thus, according to Lemma 3.11, for every initial point xi in ?+, the space shuttle can reach the EI condition in equation (4.3) by admissible controlled trajectories of the system Σsat if the satellite takes enough fuel. The periodic trajectory γxf and the associated optimal controlled trajectory with τ = τmax are illustrated in Figure 6. The profile of r and rp along optimal controlled trajectories for the DOP with τ = τmax, τmax + 100 N, and τmax 100 N, are illustrated in Figure 7.We can see from Figure 7 that the optimal controlled trajectory of the OCP for DOP with τ = τmax +100N is an admissible controlled trajectory in ? and the final point lies in ?+. While, the optimal controlled trajectory of the OCP for DOP with τ = τmax 100N cannot reach a point in ?+ by admissible controlled trajectories of the system Σ̃sat.

Fig. 6: The periodic trajectory γxf determined by the EI condition in equation (4.3) and the optimal controlled trajectory of the OCP for DOP with τ = τmax.
Fig. 7: The profile of r = r and rp along the optimal controlled trajectory of the OCP for the DOP with τ = τmax, τmax + 100 N, and τmax 100 N.

11.5Conclusion

The controllability property of the Keplerian motion around the Earth in the periodic region ? is established in this paper. According to the state constraint that the radius of the Keplerian motion has to be larger than the radius of the surface of atmosphere around the Earth, the periodic region is separated into two sets: ?+ and ?. The controlled motion in the set ?+ is the typical OTP and we obtain that the motion is controllable in the set ?+ for any positive maximum thrust. Moreover, we obtain that there exists a limiting value of τmax > 0 depending on the initial point (final point, respectively) such that the Keplerian motions for OIP (DOP, respectively) are controllable if τ > τmax. Finally, two numerical examples are simulated to show that a shooting method and a bisection method can be combined to compute the limiting value for the bound on the thrust.

11.6Appendix

In this section, we provide two sets of coordinates for points in the periodic region ? (see [8, 15] for all the results given here).

Definition 6.1 (Classical orbital elements (COE)). For x ?, define the following functions:

where 1x = [1, 0, 0]T , n = 1z × h with 1z = [0, 0, 1]T. The quantity a(x) is called the semimajor axis of the orbit γx whose shape is, thus, determined by a(x) and e(x). The angles i (x), ω(x), and Ω(x) are called the inclination of the orbit γx, the argument of perigee of the orbit γx, and the right ascension of the ascending node of the orbit γx, respectively. Then, the variables (a(x), e(x), i(x), ω(x), Ω(x), θ(x)) are called the COEs of the orbit γx.

(Note that the set of COEs is singular if e = 0 and i = 0, π.)

Definition 6.2 (Modified equinoctial orbital elements (MEOE)). For x ?, define the following functions:

where (a(x), e(x), i(x), ω(x), Ω(x), θ(x)) are the COEs defined previously. Then, the 6-tuple z = (P, ex , ey , hx , hy , l) 5 × ? gathers the so-called modified equinoctial orbit elements (MEOE), Moreover, we also have

where and W = 1 + ex cos l + ey sin l. Note that andP = h2/μ. Thus, let us define the set

Then the transformation (r, υ) : ? ?, z (r(z), υ(z)) is a covering map. Hence, ? is arc-connected if ? is.

Acknowledgment: This work was partially supported by a public grant overseen by the French National Research Agency (ANR) as part of the Investissement dAvenir program, through the iCODE Institute Project funded by IDEX Paris-Saclay, ANR-11-IDEX-0003-02.

Bibliography

[1]Baldwin, M. C.; Lu, P. Optimal Deorbit Guidance. J. Guidance Contr. Dyn. 35:93103, 2012.

[2]Bismut, J.-M. Hypoelliptic Laplacian and orbital integrals. Princeton University Press, 2011.

[3]Brand, T. J.; Brown, D. W.; Higgins, J. P. Space shuttle GNC equation document No. 124. Unified Powered Flight Guidance. NASA Rept. no. 9-10268, 1973.

[4]Bonnard, B.; Caillau, J.-B.; Dujol, R. Averaging and optimal control of elliptic Keplerian orbits with low propulsion. Systems Control Lett. 55(9):755760, 2006.

[5]Bonnard, B.; Caillau, J.-B.; Trélat, E. Geometric optimal control of elliptic Keplerian orbits. Discrete Contin. Dyn. Syst. Ser. B 5(4):929956, 2005.

[6]Curtis, H. D. Orbital mechanics for engineering students. Elsevier, 2005.

[7]Caillau, J.-B.; Noailles, J. Coplanar control of a satellite around the Earth. ESAIM Control Optim. and Calc. Var. 6:239258, 2001.

[8]Cushman, R. H.; Bates, L. M. Global aspects of integrable systems. Birkhäuser, 1997.

[9]Jean, F. Control of nonholonomic systems: From sub-Riemannian geometry to motion planning. Springer, 2014.

[10]Joosten, B. K. Descent guidance and mission planning for space shuttle. Space Shuttle Technology Conference, 113124, 1985.

[11]Jurdjevic, V. Geometric Control Theory. Cambridge University Press, 1997.

[12]Sontag, E. D. Mathematical control theory. Springer, 1998.

[13]Sussmann, H. J. Geometry and optimal control in mathematical control theory. Dedicated toRoger W. Brockett on his 60th birthday, edited by J. Baillieul and J. C. Willems. Springer, 1998.

[14]Sussmann, H. J. Résultats récents sur les courbes optimales in Quelques aspects de la théorie du contrôle. Journée Annuelle de la Société Mathématique de France, 2000.

[15]Zarrouati, O. Trajectories Spatiales. CNES Cepadues, 1987.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset